Richard Adam
Pacalaj
*a,
Yifan
Dong‡
a,
Ivan
Ramirez
b,
Roderick C. I.
MacKenzie
c,
Seyed Mehrdad
Hosseini
b,
Eva
Bittrich
d,
Julian Eliah
Heger
e,
Pascal
Kaienburg
f,
Subhrangsu
Mukherjee
g,
Jiaying
Wu§
a,
Moritz
Riede
f,
Harald
Ade
g,
Peter
Müller-Buschbaum
e,
Martin
Pfeiffer
*b and
James Robert
Durrant
*ah
aDepartment of Chemistry and Centre for Processable Electronics, Imperial College London, MSRH, 82 Wood Lane, London W12 0BZ, UK. E-mail: r.pacalaj18@imperial.ac.uk; j.durrant@imperial.ac.uk
bHeliatek GmbH, Treidlerstraße 3, 01139 Dresden, Germany. E-mail: martin.pfeiffer@heliatek.com
cDepartment of Engineering, Durham University, Lower Mount Joy, South Road, Durham DH1 3LE, UK
dLeibniz Institute of Polymer Research Dresden, Hohe Str. 6, 01069 Dresden, Germany
eTechnical University of Munich, TUM School of Natural Sciences, Department of Physics, Chair for Functional Materials, James-Franck-Str. 1, 85748 Garching, Germany
fClarendon Laboratory, Department of Physics, University of Oxford, Parks Road, Oxford OX1 3PU, UK
gDepartment of Physics, and Organic and Carbon Electronics Laboratory (ORaCEL), North Carolina State University, Raleigh, North Carolina 27695, USA
hSPECIFIC IKC, Faculty of Science and Engineering, Swansea University, Swansea SA2 7AX, UK
First published on 23rd October 2024
Vacuum-processed organic solar cells (VP-OSCs) possess many advantages for scalability. However, as the academic community focusses on high performing solution-processed OSCs, detailed studies about the relation between morphology and device characteristics in VP-OSCs are rare. Here, we present a study on a model donor/fullerene VP-OSC system deposited at different substrate temperatures. Substrate heating results in increases in current density and fill factor (FF). Changes in morphology are characterised by grazing-incidence wide-angle scattering (GIWAXS) and resonant soft X-ray scattering (RSoXS). The increase in the degree of crystallinity and preferential orientation of the donor molecule in heated samples results in enhanced absorption increasing current density. The exciton and charge separation efficiency were studied by transient absorption and photoluminescence quenching and only showed minor differences. To study the FF differences, charge transport and non-geminate recombination are studied by optoelectronic measurements and device simulations. The charge carrier kinetics are governed by a large density of trap states. While the energetic disorder and non-geminate recombination under open circuit conditions remain largely unchanged, the increased effective mobility and lower transport disorder observed in photocurrent transients explain the increased collection efficiency for heated devices. We relate this to the increased donor phase purity. Our results suggest that charge recombination and transport are governed by different aspects of disorder related to amorphous and crystalline donor phases. Quantitative comparison with high FF solution-processed OSCs reveals that the low mobility limits FF. Finally, drift-diffusion simulations give an outlook for possible performance increases through further optimisation of the deposition control.
Broader contextLab-scale solution-processed organic solar cells (SP-OSCs) now achieve power conversion efficiencies in excess of 19%. Reproducibly translating this efficiency to large area modules, i.e. transitioning from spin-coating to scalable techniques, has proven difficult. Additionally, the stability of SP-OSCs remains an open issue. In contrast to that, vacuum-processing requires higher capital investment but has been proven to be scalable in the field of organic light emitting diodes. Vacuum-processing also allows for the scalable deposition of complex multi-junction and interlayer structures. Despite these advantages, the lab-scale efficiency of vacuum-processed organic solar cells (VP-OSCs) currently lags behind that of SP-OSCs. Due to the high required capex, only a few academic groups investigate the performance limitations of VP-OSCs and how they relate to material properties and the bulk-heterojunction performance. Here, we present a study combining detailed morphological and device characterisation. The morphology was varied through substrate heating during the deposition of the active layer. Our detailed study allows us to pinpoint device properties to characteristics of the active layer morphology. Additionally, the comparison with solution-processed references helps us to identify the VP-OSC performance bottlenecks. Our findings will be useful to guide materials and morphology optimisation to further increase the performance of scalable VP-OSCs. |
The high PCEs of SP-OSCs are motivation for extensive studies of their nanomorphology and charge carrier dynamics to guide further performance enhancements. In contrast, despite their commercial potential, there are relatively few studies of the underlying factors determining the PCE of VP-OSCs.4 Herein, we focus on the correlation between nanomorphology and charge carrier dynamics in high performance BHJ VP-OSCs fabricated from a small molecule donor and C60 acceptor, and how these compare to SP-OSCs.
To date, the efficiency advances achieved with NFAs in SP-OSCs have not been mirrored in VP-OSCs. The highest performing single-junction VP-OSCs still rely on fullerene acceptors.16 In part, this is due to the molecular size constraint imposed by thermal evaporation prohibiting the use of similarly large conjugated ring structures as present in the highest performing NFAs.4,17 This difference in molecular size also plays a role in the aggregation behaviour of high performing solution vs. vacuum-processed BHJs and thereby impacts performance.18
High performance SP-OSC systems typically exhibit relatively ordered domains, with a composition of the mixed polymer-rich domains near the percolation threshold for electron transport and prevention of over-purification and excessive aggregation.18–22 This has been shown to improve the charge generation and collection efficiency. Phase separation and purity can be aided through the use of solvent additives or post-deposition thermal/solvent-vapour annealing.23–26 Evaporable small molecule donors paired with fullerene acceptors are believed to form intermixed blends, related to their similar size and stronger intermolecular interactions between donor and acceptor molecules.18,27 Limited morphological optimisation for VP-OSCs28 may be achieved through the use of templating substrates,27,29 co-evaporants inducing crystallization,30–33 thermal annealing34,35 or by altering the substrate temperature during evaporation.34,36–45
It has previously been shown that vacuum-processed blends exhibit lower fill factors (FF) than efficient SP-OSCs.17 This may originate from effects of field-dependent generation or charge collection. The former depends on the intrinsic energetic offset and excited state lifetimes as well as the morphology dependent contributions to these. Empirically, the latter has been inferred from the drop in FF typically observed for active layers thicker than 50 nm, thus limiting optical absorption and hence short-circuit current density (JSC). In contrast, SP-OSCs typically achieve the highest performance at active layer thicknesses ≈100 nm,46,47 indeed for some devices >300 nm.20,48–52 A key challenge for VP-OSC is therefore gaining a better understanding of the interplay between deposition parameters, active layer morphology and the efficiency of charge separation, and collection (transport vs. non-geminate recombination) to maximise JSC and FF.
Here, we present a study on the interplay between morphology, charge carrier dynamics and device performance for a VP-OSC model system: DCV-V-Fu-Ind-Fu-V:C60 (2:1)53 that is structurally related to commercially relevant systems. The active layer morphology is altered by substrate heating during evaporation. This was empirically shown to improve the device performance in low performing planar34,37,39–42 and bulk heterojunctions.36,38,43–45 Few studies link the BHJ performance with film morphology (improved crystalline order and molecular orientation) but do not provide detailed device characterisation.34,36 In this study we go beyond previous literature by combining detailed morphological and device characterisation in a relevant model blend.
We employ grazing-incidence wide-angle X-ray scattering (GIWAXS) and resonant soft X-ray scattering (RSoXS) to study substrate heating induced differences in molecular orientation and crystallinity, and the size and purity of phase-separated domains, respectively. Subsequently, we analyse the impact of the changed active layer morphology on solar cell operation – from charge generation (absorption, exciton quenching efficiency, charge separation) to charge collection (non-geminate recombination vs. charge extraction). Active layer films were characterised using UV/Vis absorption, photoluminescence quenching and ultra-fast transient absorption spectroscopy. To address the collection efficiency limitations, we employed transient photovoltage/photocurrent measurements and charge extraction analyses. We focus in particular on the impact of the energetic disorder in the blends that has been shown to strongly impact the FF of VP-OSCs.54 Our analyses of the collection efficiency are compared to two SP-OSC references – employing small-molecule:fullerene (BTR:PC70BM) and polymer:NFA (PM6:Y6) photoactive layers. Both of these solution processed references are known for their high collection efficiencies and thickness tolerance.20,55–57
We show that the increased crystallinity in films deposited on heated substrates explains the observed increases in photocurrent. While the non-geminate recombination is largely unaffected by the change in morphology, the increased effective mobility in the elevated temperature samples leads to an improved collection efficiency and thereby fill factor. The comparison with the solution-processed references reveals a strong carrier density dependence of the mobility and recombination rate constant in the VP-OSCs.58–61 While recombination kinetics are comparatively slow in the VP-OSCs, the effective mobility is found to be the bottleneck compared to the SP-OSC references. Combining optoelectronic measurements and device simulations using the software package OghmaNano62,63 we assign this to slow hole transport. Our morphological data suggests that substrate heating can partially mitigate this by increasing the donor crystallinity and phase purity leading to an increase in the effective mobility. We conclude by providing an outlook for possible performance increases for single junction VP-OSCs of DCV-V-Fu-Ind-Fu-V:C60 that could be achievable by further optimisation of the deposition control and/or adjustments of the molecular structure yielding improved donor aggregation.
The active layer thickness of 50 nm was chosen as a compromise between absorption and charge carrier collection efficiency. While the optimum performance is achieved for thinner active layers, 50 nm also helps to reduce the impact of the device capacitance on optoelectronic measurements, as discussed in the methods section (see ESI†). The active layer is evaporated on a thin layer of undoped C60 and capped with undoped BPAPF (see ESI,† Fig. S1 for the structure), the electron and hole selective contacts, respectively. This interlayer structure minimises the effect of surface recombination of minority carriers at the contacts. An additional layer of n-doped C60 and p-doped BPAPF complete the respective interlayers. These aid the reduction of the series resistance by making the bulk of the transport layers more conductive.65
Fig. 1(b) shows a visual representation of the energy alignment of the device. Note that the offsets at the interfaces are not to scale. Further details about the device preparation can be found in the (ESI†). Fig. 1(c) shows the JV-curve of the RT and 50 °C device. Heating the substrate during evaporation yields an improved FF (61 vs. 53%) and JSC (15.1 vs. 12.9 mA cm−2) resulting in an improvement of the PCE from 5.8% to 7.9%. The VOC of the 50 °C device only shows a small increase of about 10 mV compared to the RT device. This together with the very similar dark JV-curves shown in ESI,† Fig. S2 indicate similar voltage losses in the system.
ESI,† Fig. S3 shows the JV-curves of the SP-OSC references BTR:PCBM and PM6:Y6. The comparison illustrates the FF limitations of the VP-OSCs. The origin of the FF limitation, i.e., field-dependent generation or charge collection, will be investigated in more detail below.
At reverse and low forward bias, both VP-OSCs show very little variation in the normalised JV-curve shape with increasing light intensity. This is indicative of the JV-curves being dominated by a first order process rather than intensity dependent charge collection in this region. Efficient charge collection at JSC up to 1 sun equivalent light irradiation is also demonstrated by the excellent JSC linearity in Fig. 2(f). At light intensities above 1 sun, differences become discernible with a pronounced decrease in JSC linearity of the VP-OSCs. The onset of the decrease occurs at lower light intensity for the RT vs. the 50 °C device (approximately 1 sun vs. 2 sun, respectively) in line with a reduced collection efficiency.
Given the good linearity up to 1 sun, the reduced JSC in the RT device compared to that deposited at 50 °C (Fig. 1(c)) is more likely related to differences in absorption and field dependent geminate losses, as assessed below. The current saturation (ratio of JSC to reverse bias current) of the 50 °C device around JSC is improved and comparable with that of the non-fullerene reference, while the large offset fullerene reference (BTR:PCBM) shows the highest current saturation at JSC.
These results indicate that the morphological modifications achieved by varying the substrate temperature during the evaporation process improve both the charge generation and collection elements of device function. While the charge collection is significantly improved in the 50 °C device, it still constitutes a performance bottleneck compared to the SP-OSC references. Following the in-depth assessment of the active layer morphology, we will assess the charge generation, recombination and extraction in more detail.
Fig. 3 (a) and (b) 2D-GIWAXS diffraction patterns for RT and 50 °C blend films, respectively. The sharp reflection in (a) is due to specular reflection from the edged Si substrate (see ESI,† Fig. S4). (c) 1D diffraction patterns obtained from out-of-plane cake cuts of the RT and 50 °C film, respectively. The Gaussian fits are shown in solid lines together with the corresponding FWHM values used to estimate the coherence length in the main text. (d) RSoXS data at resonance on a log–log scale for samples deposited at different substrate temperatures. The table indicates the long period L and integrated scattering intensity ISI normalized to the 60 °C sample. |
The full width at half maximum (FWHM) of the reflexes is a measure of the coherence length of the respective lattice planes (hkl), which is related to the crystallite size according to the Scherrer equation, where the size Lhkl ∼ 2π/FWHM.69 Additional broadening effects due to the paracrystalline nature and lattice disorder of soft matter make a defined calculation of crystallite sizes non-trivial,70 however, a qualitative estimation of their lower limit is possible. In ESI,† Fig. S4(c), the FWHM of the (010) reflex remains constant at (0.38 ± 0.01) Å−1, while the FWHM of the (100) reflex narrows from (0.23 ± 0.01) Å−1 to (0.15 ± 0.01) Å−1 upon increasing the substrate temperature. This evolution suggests that the minimum crystallite size along the [010] direction remains constant, while it is 1.5 times larger along the [100] direction in the 50 °C film.
Pole figures of radially integrated azimuthal tube cuts along the (100) reflex (ESI,† Fig. S4(d)) and the (010) reflex (ESI,† Fig. S4(e)) reveal the material quantity of crystallites in preferred orientation with respect to the substrate.71 For both substrate temperatures, face-on and edge-on populations are observed. The results show a strongly increased preferential orientation for the (100) reflex in the 50 °C film judged by relative changes in scattering intensity, with about 3 times more crystallites in face-on orientation and about 8 times more in edge-on orientation, as compared to the RT film. When comparing the different preferential orientations for the (100) reflex, the amount of face-on orientation is about 12 times higher than edge-on orientation in the RT film, and about 5 times higher in the 50 °C film. The (010) reflex shows an about 1.3 times higher face-on orientation in the 50 °C film compared to the RT film. The increase in the (100) out-of-plane reflection indicative of an overall higher crystallinity is likely to increase the fraction of the integral transition dipole moment (expected to be parallel with the long molecular axis) within the substrate plane for the 50 °C film, thus enhancing the absorption of incident visible light within the photoactive blend, as we discuss further below.
Besides the average degree of crystallinity, the size of phase-separated domains and their purity are critical for BHJ device operation. Resonant Soft X-ray Scattering (RSoXS) is a technique capable of measuring the relative composition of domains as well as quantifying the size of the domains. The technique has previously been applied to SP-OSCs.72–74Fig. 3(d) shows blends deposited at different substrate temperatures (RT, 40 °C, 60 °C). The two slightly differing heated conditions were chosen in order to test the robustness of the established trend. Additionally, the donor content is slightly reduced to 60% from 66.7% (by weight).
The presence of a strong peak at q ∼ 0.3 nm−1 shows that phase separation occurs already at room temperature. With increasing substrate temperature, the long period L – indicative of the domain size – continuously increases from 16.7 nm at RT to 20.3 nm at 60 °C. These values are smaller than those typically obtained for SP-OSCs using the same technique.26,75,76 The increase in domain size upon heating the substrate coincides with increased domain purity. This can be seen qualitatively by the higher RSoXS scattering intensity for higher temperature samples. When quantifying the relative domain composition with the relative integrated scattering intensity ISI,73 the RT ISI is a third of the 60 °C sample. Note, that to be as close as possible to the device stack, the blends are sandwiched between layers of C60 and BPAPF which makes a quantitative interpretation of the ISI challenging. However, trends and the qualitative behaviour of the blends upon annealing can be deduced from the data as discussed alongside further details of the RSoXS measurements in ESI,† Fig. S5 and S6.
In summary, the RT sample already shows some degree of crystalline order and phase separation. Two populations of crystallites with edge-on and face-on orientation are observed. For the 50 °C sample, we observe an overall higher degree of crystallinity, tighter packing of molecules, larger crystallites in the lamellar direction and phase-separated domains becoming larger and purer. The crystallite size or disorder in π–π direction seems unchanged. Both orientation populations are also observed at 50 °C but the face-on orientation becomes more prevalent.
These observations agree with the GIWAXS data discussed above. The higher degree of order observed in the 50 °C film (increased crystallinity of edge-on and face-on phases) indicates that fewer molecules are tilted upwards towards an end-on orientation, which is consistent with the increased optical absorption. Interestingly, the change in donor aggregation determined from the GIWAXS data does not lead to a significant shift in absorption, as observed in related DCV based small molecule donors (also see EQE absorption onset in ESI,† Fig. S7).29,79 This would be in line with the presence of crystalline albeit smaller donor domains even in the RT films as also suggested by the morphological data.
The external quantum efficiency (EQE) data (Fig. 4(b)) indicate that the increase in film absorption observed for the 50 °C films is also reflected in an increase in photocurrent generation in complete devices. Measurements of the field dependent EQE were used to determine the quantum efficiency of charge separation by calculating the ratio of the EQE determined at 0 V and −3 V (Fig. 4(b)). It indicates a small improvement in the efficiency of charge separation for the 50 °C device, especially in the C60 absorption region. This is in line with the slightly improved saturation at reverse bias, as discussed above for the light intensity dependent JV-curves, most likely stemming from a small reduction in field dependent geminate recombination.
To assess the efficiency of exciton separation, the photoluminescence quenching efficiency (PLQE) of the BHJs was measured, as is commonly done for fullerene-based devices. Fig. 4(c) shows that the absorption corrected donor PL intensity is reduced by 86% and 96% in the 50 °C and RT blends respectively, relative to the neat donor film. Additionally, the exciton decay lifetime upon optically pumping the donor was measured via ultrafast transient absorption spectroscopy (see Fig. 5). The time resolved measurements show a reduction of the exciton lifetime (photo-induced absorption signal around 900 to 920 nm, see Fig. 5(a) and (b)) from ∼50 ps in the neat donor film to 2.9 ps and 1.9 ps for the 50 °C and RT films, respectively (see kinetic Fig. 5(c)). The fluence dependent kinetics of the blends as well as the neat donor lifetime and TAS spectra are available in ESI,† Fig. S8(a)–(f), respectively. Note, that the neat donor exciton lifetime was extracted as a weighted average of a biexponential decay with a significant slow decay phase. The reported value is significantly shorter than previously reported values for NFA materials.80
These exciton decay times are in good agreement with the observed trend of the PLQE. The lower PLQE and slower exciton decay observed for the 50 °C films are indicative of a less intermixed morphology, increasing the distance excitons must diffuse to reach a donor/acceptor interface,81 in agreement with the GIWAXS (more crystalline) and RSoXS (larger and more pure domains) data discussed above. However, the effect of reduced exciton separation efficiency on current generation is compensated by the enhanced light absorption of the 50 °C device, consistent with its overall higher JSC.
Fig. 6 (a) Differential capacitance vs. open circuit voltage (as measured and corrected for geometric capacitance) together with the respective fits for the 50 °C device. See ESI,† Fig. S9 for RT data. (b) Charge carrier density vs. the open circuit voltage referenced to the open circuit voltage at 2 sun equivalent illumination to enable a better comparison of the slopes. Exponential fits of the form n0exp(qVoc/2Ech) and the bandtail of an ideal semiconductor are shown. (c) Effective mobility and effective bimolecular recombination coefficient vs. charge carrier density. Both axes scaled proportionally to each other to reveal the proportionality of the charge carrier dependence of mobility and recombination coefficient. (d) Normalised photocurrent transients measured using variable load resistances as described previously.48,77 The dashed lines were simulated using OghmaNano using a global fit to experimental data as described below. |
Fig. 6(b) shows the spatially averaged charge carrier density of the VP-OSCs and reference SP-OSCs plotted vs. the VOC used as a proxy of the quasi-Fermi level splitting (QFLS). For better comparison of the slope of the tail state density against the solution processed references, the x-axis was offset by the VOC at 2 sun equivalent illumination. Additionally, the ideal semiconductor charge density, with Ech equal to the thermal energy, is shown.87 The charge carrier densities and fitted Ech values only show minor differences between the RT and 50 °C device. Characteristic energies, 53 and 50 meV for the RT and 50 °C devices respectively, are typical of many SP-OSCs56,81 but significantly larger than the values obtained for BTR:PCBM (37 meV) and PM6:Y6 (31 meV). For BTR:PCBM and PM6:Y6, the low Ech has been linked with their abilities to support efficient charge collection for thicker photoactive layers. Similar trends in CT state disorder have previously been observed in sensitive EQE measurements comparing PM6:Y6 and a range of VP-OSCs88 while transport based techniques such as temperature dependent SCLC have yielded significantly different trends for the polaron disorder.89 Below, high-resolution photocurrent transient and simulation results will be discussed that similarly yield different values of Ech.
Transient photovoltage (TPV) was employed to determine the charge carrier recombination kinetics at the same open circuit and light bias conditions (up to 6 sun) as the differential capacitance results presented above. Based on the lifetimes τ and charge carrier densities n (see ESI,† Fig. S10 for τ vs. n), an effective bimolecular recombination coefficient can be calculated as kBI = (nτ)−1.90 The right axis of Fig. 6(c) shows the effective bimolecular recombination coefficient vs. the charge carrier density for the four devices studied. The 50 °C and RT device exhibit similar kBI, with a slightly increased recombination rate observed for the 50 °C device. For the charge carrier density range characterised herein, both the VP-OSCs and the BTR:PCBM reference exhibit significantly slower recombination compared to PM6:Y6. For the SP-OSC references, the values are in excellent agreement with those determined by other methods such as bias assisted charge extraction (BACE) or photoinduced absorption (PIA).89
The solution-processed reference solar cells exhibit near ideal second order recombination kinetics as illustrated by the fitting of the lifetime vs. carrier density (see ESI,† Fig. S10) and the constant bimolecular recombination coefficient (Fig. 6(c)). The VP-OSC devices exhibit non-ideal (apparent third order) recombination and consequently a charge carrier density dependent kBI, attributed to their higher energetic disorder.86 To verify the validity of the analyses, the VOC was reconstructed using the lifetime and carrier density data as well as JSC as a proxy for the generation current. An excellent agreement is obtained between the measured and reconstructed VOC, as shown in ESI,† Fig. S11, highlighting the reliability of the charge carrier density and lifetime data.
Following the assessment of the recombination kinetics, we turn to measurements of the effective charge carrier mobility (μeff). The effective mobility was assessed by measuring the charge carrier density accumulated in the device under short circuit conditions with varying background light intensities.91 This method yields a charge carrier density dependent effective mobility dominated by the slower carrier type that has been shown to determine the electronic quality and FF.92 This method was chosen as it can be performed on devices under operating conditions (without the need for different interlayers as in single carrier devices) and under realistic charge carrier densities.93,94 The resulting values of the charge carrier density dependent mobility are plotted on the left axis of Fig. 6(c). The mobility of the 50 °C device shows an improvement by a factor of two relative to the RT device across the whole range of the measured charge carrier density. Given their similar recombination coefficients, this increase in mobility explains the increased FF of the 50 °C device, as we discuss further below. As both y-axes are plotted proportionally to each other, it is possible to directly compare the charge carrier dependence of both μeff and the kBI. Both VP-OSCs show a strong charge carrier dependence in μeff similar to the kBI at low charge carrier densities. BTR:PCBM shows only a weak carrier density dependence while PM6:Y6 shows virtually no carrier density dependence of the mobility, consistent with the lower energetic disorder and the absence of trap filling effects.
ESI,† Fig. S12 shows normalised photocurrent transients; these were used to determine the charge density in the devices at short circuit. The faster current transients of the SP-OSC devices qualitatively indicate a higher mobility in these blends (ESI,† Fig. S12(e)), consistent with Fig. 6(c). The strong biphasic character observed for the VP-OSC devices has previously been related to slow extraction from deeper trap states in evaporated bulk heterojunctions, most pronounced at low light intensity (see ESI,† Fig. S12).35 High resolution transient photocurrent measurements have previously also been used to quantitively assess the characteristic energy.62,95,96Fig. 6(d) shows the high resolution current transients measured with variable load resistances for the RT and 50 °C devices. In contrast with the energetic disorder derived from the carrier density vs. VOC, these transport-based measurements combined with a global fit of the optoelectronic data using drift-diffusion simulations (OghmaNano) indicate higher values of Ech. A global fit of all data results in values of 121 meV and 89 meV for the RT and 50 °C devices, respectively, contrasting the lower values of characteristic energy determined from measurements at open circuit (53 meV and 50 meV, respectively). We therefore assign the improved FF in the 50 °C device to the overall increased effective mobility and a reduction in the transport related disorder as observed in these current transients. At the same time, the non-geminate recombination kinetics are similar for both VP-OSCs and comparatively slow compared to the SP-OSC references. The different magnitudes of energetic disorder from our open circuit charge extraction measurements (Fig. 6(b)) and short circuit photocurrent transient (Fig. 6(d)) analyses are discussed below.
Both VP-OSCs studied herein exhibit an absorption onset around 800 nm (Eg = 1.72 eV) together with a VOC of around 0.85 V. This voltage loss compares unfavourably with PM6:Y6 which exhibits a similar VOC at a much smaller optical bandgap of 930 nm (Eg = 1.33 eV). It is however more comparable to typical voltage losses observed for SP-OSCs employing fullerene based acceptors including BTR:PCBM (Eg = 1.92, VOC = 0.90 V), attributed at least in part to a larger energetic offset requirement for efficient exciton separation with fullerene acceptors.16,101–103 Besides this limitation posed by the persistent use of fullerene acceptors in high performance VP-OSCs, the main performance limitation arises from the poor charge collection efficiency limiting the device FF even in 50 nm thin devices as illustrated in Fig. 2.
Table 1 lists the values of the effective mobility, calculated Langevin recombination coefficient, experimental bimolecular recombination coefficient as well as the Langevin reduction factor and electronic quality Q determined at a charge carrier density corresponding to MPP under 1 sun equivalent illumination (see the caption of ESI,† Fig. S13 for details about the calculation). Additionally, QJV was calculated directly from the JV-data using the equation provided by Kaienburg et al. reproducing the trends of the experimental values. The value of Q obtained for the RT device is comparable to those reported for similar donor molecules such as the DCV5T-Me:C60 blend.17 While it is improved 3–4 fold in the 50 °C device to values of 0.53 and 0.39 for QExp and QJV, respectively, more crystalline blends like ZnPc:C60 achieve values in excess of 1.17 The values determined herein for the SP-OSCs are within the range reported previously. BTR:PCBM stands out, highlighting the superior electronic quality of some fullerene-based devices even compared to high performing NFAs.17,92
Material | μ eff [cm2 V−1 s−1] | k L [cm3 s−1] | k BI [cm3 s−1] | γ | Q Exp [cm1.6 V−2 s1.2] | Q JV [cm1.6 V−2 s1.2] |
---|---|---|---|---|---|---|
DCV-V-Fu-Ind-Fu-V:C60 (RT) | 1.2 × 10−5 | 6.1 × 10−12 | 4.3 × 10−12 | 1.4 | 0.18 | 0.10 |
DCV-V-Fu-Ind-Fu-V:C60 (50 °C) | 2.0 × 10−5 | 1.1 × 10−11 | 4.1 × 10−12 | 2.6 | 0.53 | 0.39 |
BTR:PCBM | 3.5 × 10−4 | 1.8 × 10−10 | 4.5 × 10−12 | 40 | 142 | 819 |
PM6:Y6 | 1.2 × 10−4 | 6.4 × 10−11 | 2.3 × 10−11 | 2.8 | 5.0 | 23.3 |
Since the mobility and bimolecular recombination coefficient are strongly charge density dependent in the evaporated cells, Q, kL and γ also vary with light intensity (see ESI,† Fig. S13 for the light intensity dependence of QExp, QJV and γ). Comparison with a single value of γ, as often quoted in the literature, may therefore be misleading, especially if it is based on a value of kL calculated from SCLC mobilities determined at higher charge carrier densities unrepresentative of 1 sun operating conditions. It is evident from Table 1 that the Langevin reduction factors of the PM6:Y6 and the 50 °C devices are similar at 1 sun, and therefore that this Langevin based analysis is insufficient to explain the greater charge collection and FF limitations of the 50 °C device. By giving a higher weighting to the mobility, the electronic quality QExp proves to be a more suitable measure to explain the FF variations. The similarity between the values obtained from transient optoelectronic experiments and those directly calculated from the JVs highlight the potential of QJV as a screening tool.
The presence of energetic disorder can often lead to discrepancies in experimentally determined charge carrier lifetime data, depending on the charge carrier density regime these were measured at.105 To illustrate this, we used the fluence dependent decay of the polaron signal in the transient absorption measurements (see ESI,† Fig. S8) to calculate the kBI. An estimate of the charge carrier density excited by the laser pulse was calculated from the device EQE at the excitation wavelength. Fig. 7 shows the results of the calculation side by side with the values obtained by the combination of TPV and DC measurements.
It is apparent that the optoelectronic measurements are conducted at light intensities relevant to solar irradiation conditions, while the pump pulses used for transient absorption measurements generate orders of magnitude higher charge carrier densities. In the case of charge carrier density dependent bimolecular recombination coefficients – as observed previously for fullerene based SP-OSCs84,106,107 as well as the VP-OSCs studied herein – this will lead to discrepancies between high-fluence TAS and comparatively low light intensity optoelectronic measurements. Here, it allows us to show that the kBI saturates above a carrier density of 1018 cm−3 for the VP-OSCs. Such a behaviour was previously reported in the literature for fullerene based SP-OSCs and may be related to the saturation of trap states and the more dominant role of free-to-free recombination.106 While the recombination kinetics at irradiation levels relevant to device operation were similar, the 50 °C device shows a two-fold increased saturated kBI of about 6 × 1011 cm3 s−1. Both saturated values exceed the value obtained for the PM6:Y6 reference. It is also striking that for PM6:Y6, we obtain good agreement between the kBI determined from TAS and TPV/CE studies, consistent with greatly suppressed energetic disorder in this device and near ideal bimolecular kinetics.
To investigate the contribution of the different recombination channels in more detail as well as shed light on the origin of the slow phase observed in current transients, we employed the simulation software OghmaNano. The software treats intra-gap trap states as an ensemble of trap states with discrete energy levels making up an exponential tail and solving the trapping and recombination rate equations for every trap level independently.62 The simulation parameters shown in ESI,† Table S1 were extracted by performing a global fit to light and dark JV-curves, the 1 sun TPV transient, high resolution photocurrent transients at variable bias (−1 V, 0 V, 0.5 V) as well as VOC, JSC and n vs. light intensity (see ESI,† Fig. S14 and S15).
The charge carrier lifetimes and effective mobility were extracted from the simulations and compared to the values determined experimentally (see ESI,† Fig. S16). The good agreement in the general trends and the small deviation in absolute values within a factor of two indicate the overall applicability of the model to the devices studied herein. We note that, no free-electron-to-free-hole recombination channel was necessary to reproduce the optoelectronic experimental results. However, at the much higher fluences experienced by the device in a TAS experiment, an additional direct free-to-free recombination channel (kfree-to-free of 6.6 × 10−11 and 3.3 × 10−11 cm3 s−1 for 50 °C and RT, respectively) was needed to reproduce the higher light intensity trends (see dashed lines in Fig. 7) in line with a pronounced contribution of free carrier recombination.
While the experimental data of the 50 °C device can be accurately reproduced using symmetric trap populations (symmetric hole and electron Ech and trap densities), the global fit to the RT data required an asymmetry of the trap density and larger Ech (121 meV vs. 89 meV for RT and 50 °C, respectively). These values of the characteristic energy that are primarily determined from the fitting of the photocurrent transients differ significantly from the values obtained for the combined tailslopes determined from charge density measurements vs. VOC (53 meV and 50 meV for RT and 50 °C, respectively). The latter together with lifetime data from TPV was used to accurately reconstruct the experimentally determined VOC including their ideality, as shown in ESI,† Fig. S11. This indicates that the disorder obtained from charge extraction measurements determines non-geminate recombination at open circuit conditions (n > 1015 cm−3) while the photocurrent transients may be governed by energetically lower lying traps and asymmetric populations.
Potentially in line with this observation, Saladina et al. recently reported a power law DoS in DCV-V-Fu-Ind-Fu-V:C60. They inferred this from light and temperature dependent ideality factors.53 A consequence of their observations would be a narrower DoS compared to an exponential bandtail but with an increasing breadth of states (increasing Ech) deeper into the bandgap. Qualitatively this may be similar to our observations of a larger Ech fitted to the slow phase of current transients corresponding to states deeper in the gap and asymmetric bandtails. Another possible explanation for asymmetric transport properties affecting the photocurrent transients may be the increased presence of spatially isolated donor domains in the RT film as supported by the lower phase purity and smaller domain size shown in RSoXS and GIWAXS data, respectively. While contributing to non-geminate recombination, these domains may slow down charge extraction as holes have to travel through amorphous and intermixed regions leading to the observed disorder in current transients and the asymmetry of the hole and electron density that was necessary to achieve a good global fit. In support of this observation, ESI,† Fig. S17 illustrates that a reduction in donor content leading to less aggregated donor phases results in a similarly pronounced slow phase of the current transient as observed in the RT device. Charge transport between isolated donor domains via tunnelling or hole transport through C60 has previously been linked to slow current transients.108 The polaron disorder in amorphous phases has previously been established as a key determinant of hole mobility in small molecule hole transport materials.109,110
Our observations may also point to fundamental discrepancies in the determination of disorder through transport-based techniques (current transients, temperature dependent SCLC), techniques relating charge accumulation and QFLS (impedance spectroscopy, charge extraction vs. VOC, VOC ideality), and techniques based on the fitting of CT state emission/absorption. These have previously yielded a wide range of disorder parameters and even differ in the DoS shape (Gaussian vs. exponential vs. power law).53–56,87–89 A full discussion of the different techniques and results is however beyond the scope of this study.
In summary, charge carrier build-up and recombination under open circuit conditions are largely unaffected by the change in substrate temperature during deposition. Both are characterised by a broad distribution of tailstates compared to PM6:Y6 and BTR:PCBM but comparable to other SP-OSC materials as measured by the same methodology. Simulation results consistently explain the observed charge carrier density dependence of the effective recombination coefficient and mobility through a recombination formalism involving trapped and free carriers. Our results show that while the non-geminate recombination rate at operational charge carrier densities is sufficiently slow compared to the SP-OSC references (∼4 × 10−12 cm3 s−1), the observed low collection efficiency and FF can be rationalized by the overall low mobility (∼10−5 cm2 V−1 s−1), likely resulting from large energetic disorder during charge transport. Depositing at higher substrate temperatures induces favourable changes in microstructure, improving structural order and phase separation, which goes along with improved energetic disorder for charge transport and a twofold increase in effective mobility. In the following, we explore the gains in single-junction device performance that would result from further improving morphology, disorder and charge carrier dynamics.
Fig. 8 (a) Simulated fitted JV-curves of the RT and 50 °C device. Additionally shown are the JV-curves for the RT device with iteratively changing the generation current (Jgen), the trap symmetry and density (ntrap), and the mobility of carriers at the band edge (μfree) and the capture cross-section of the charge carriers (rcapture) to the values obtained from the global fit of the 50 °C device. (b) Simulated JV-curves visualising the effect of adjusting Ech, free mobility, trap density and active layer thickness to achieve 10% PCE. (c) PCE corresponding to the simulated JV-curves in (a) and (b). Improvements beyond the performance of the 50 °C device parameter changes described in the legend of (b). The full list of modelling parameters can be found in ESI,† Table S1. |
Further reducing transport Ech to 50 meV yields a large improvement in FF.54 Comparatively, the influence of halving the trap density and an increased free mobility yields only small further improvements (see ESI,† Table S1 for the full list of simulation parameters). This reduction in the energetic disorder improves the collection efficiency and allows for the active layer thickness to be increased to 70 nm. At 70 nm the absorption of the active layer reaches its first interference maximum leading to an improved JSC outweighing the reduction in FF (see ESI,† Fig. S18 for transfer matrix simulations). Note that in VP-OSCs the inclusion of doped transport layers as optical spacers can shift the interference maxima and further improve absorption.17,111,114 This demonstrates the potential that lies in improving the collection efficiency through morphological and structural modifications of existing high-performance systems.
To close the efficiency gap with SP-OSCs, single-junction VP-OSCs will furthermore require a reduction in the voltage losses related to the D/A energy offset required for efficient charge generation as demonstrated for modern NFAs in SP-OSCs. Achieving high generation efficiency at low offsets may also require an increased exciton lifetime of the donor materials.80,115 Kaienburg et al. demonstrated that in principle evaporable donor molecules may achieve reduced voltage losses when paired with the NFA Y6 in hybrid devices where the donor is evaporated onto the solution-processed NFA.16 While the introduction of evaporable NFAs may allow for better energy level tuning, it remains to be seen whether charge generation properties of solution-processed acceptors like Y6 can be mirrored in smaller molecules suitable for evaporation. Additionally, our results show that despite the lower C60 content by weight, the electron mobility is not limiting performance. The use of evaporable NFAs may have detrimental effects on charge collection due to the anisotropy of planar molecules and likely intermixed blend morphologies. Therefore, potential high-performance evaporable NFAs should exhibit high mobilities and sufficiently phase separated blends to ensure good charge transport through the respective crystalline phases. A particular focus should therefore lie on optimising the interaction parameters and miscibility of the similarly sized evaporable donor and NFA molecules116,117 as well as the optimisation of deposition parameters.
Substrate heating also led to a substantial increase in FF. To investigate this, we determined the non-geminate recombination rate constants and effective mobility. Both are highly charge density dependent which is characteristic of charge carrier dynamics in the presence of a high trap density. The origin of the FF improvement lies in the two-fold increase of the effective mobility related to the reduction in energetic disorder and deep lying tailstates inhibiting charge transport. At the same time, recombination kinetics remained largely unaffected under operational conditions. The reduced mobility and the slow phase in current transients are tentatively assigned to amorphous/spatially isolated donor phases that trap holes. This is further supported by the RSoXS measurements that indicate a higher phase purity for films deposited at elevated substrate temperatures.
Our morphological data together with the optical and optoelectronic characterisations suggest that the phase purity and the decrease in amorphous phase size are key to reducing the disorder related to charge transport that limits the device FF. Further increasing the substrate temperature during evaporation may increase the size of crystalline domains leading to further reduction of the exciton quenching efficiency or cause changes in the molecular orientation resulting in a loss in absorption.110 Due to the large anisotropy of the small molecule blends, our results demonstrate that morphological optimisation of evaporated organic solar cells is non-trivial but may enable large performance increases.
Simulations revealed that the reduction in intra-gap states, without negatively affecting other aspects of device performance, could lead to further significant performance increases in this evaporated model system beyond 10% PCE for single junction devices. This requires careful morphological optimisation and may further be aided by modifications of the molecular structure that could aid aggregation e.g. steric groups that have been shown to affect morphology in blends.64
Further improvements in single-junction efficiency will require a reduction of the voltage losses resulting from the trade-off between the energy level offset and efficient charge generation by addressing the relatively short exciton lifetime.80 Evaporable NFAs may provide an avenue to reducing voltage losses.16 When replacing fullerenes with NFAs, good electron mobility and sufficient phase separation in blends with similarly sized donors will be key to achieving good collection efficiency and overall performance.
Boosting the single-junction efficiency of vacuum-processed solar cells will translate to improved multi-junction efficiencies and further raise the prospects of their large-scale commercialisation.
Footnotes |
† Electronic supplementary information (ESI) available: Including experimental details and additional supplementary results. See DOI: https://doi.org/10.1039/d4ee03623a |
‡ Current address: Chemistry and Nanoscience Center, National Renewable Energy Laboratory, 15013 Denver West Parkway, Golden, CO 80401, USA. |
§ Current address: Department of Chemical and Biological Engineering, Hong Kong University of Science and Technology, Room 6542, Academic Building, Clear Water Bay, Kowloon, Hong Kong. |
This journal is © The Royal Society of Chemistry 2024 |